Chapter 4

The Fundamental Laws of Thermomechanics

Abstract

This chapter discusses the kinetics and thermodynamics of a continuum, introducing concepts such as force, moment, momentum, stress, energy, work, heat, and entropy. Special attention is devoted to developing the fundamental laws (or first principles) of thermomechanics. These include the mechanical conservation laws of mass, linear momentum, and angular momentum, as well as the first law of thermodynamics (or conservation of energy). Each of the fundamental laws is first postulated as a primitive statement (in words), from which we carefully progress to material, integral, and pointwise forms. Both Eulerian (present configuration) and Lagrangian (reference configuration) representations of the fundamental laws are discussed. The chapter concludes with a presentation of the second law of thermodynamics in the form of the Clausius-Duhem inequality.

Keywords

kinetics

thermodynamics

conservation laws

mass

momentum

energy

stress

4.1 Mass

Consider again the configuration of body Bsi169_e at time t (i.e., the present configuration) in which Bsi170_e occupies the region Rsi171_e of E3si172_e bounded by the closed surface Rsi173_e; a part Ssi174_e of body Bsi175_e occupies a region PRsi176_e bounded by a closed surface Psi177_e; refer to Figure 3.9. Each part Ssi178_e of body Bsi179_e at each instant of time is assumed to be endowed with a nonnegative measure M(S,t)si180_e, called the mass of part Ssi181_e.

We define the mass density ρ at particle Y in the present configuration by

ρ=ˉρ(Y,t)=limV0M(S,t)V,

si182_e  (4.1)

where V=V(P)si183_e is the volume of the region Psi184_e occupied by subset Ssi185_e as it collapses to particle Y. Since V is positive and M(S,t)si186_e is nonnegative, density ρ is nonnegative.

In continuum mechanics, it is assumed that the measure M(S,t)si187_e is absolutely continuous, i.e., in every configuration a part Ssi188_e occupying a sufficiently small volume has arbitrarily small mass M(S,t)si189_e. Thus, concentrated point, line, and surface masses are excluded, and the limit (4.1) always exists.

The mass of part Ssi190_e of body Bsi191_e and the mass of body Bsi192_e itself at time t can be expressed in terms of density ρ by

M(S,t)=Pρdv,M(B,t)=Rρdv,

si193_e  (4.2)

where dv is an element of volume in the present configuration (refer to Section 3.6).

We define the mass density ρR at particle Y in the reference configuration by

ρR=ˉρR(Y)=limVR0M(SR)VR,

si194_e  (4.3)

where VR=VR(PR)si195_e is the volume of region PRsi196_e occupied by subset Ssi197_e in the reference configuration, and M(SR)si198_e is the mass of Ssi199_e in the reference configuration, as Ssi200_e collapses to particle Y.

The mass of part Ssi201_e of body Bsi202_e and the mass of body Bsi203_e itself in the reference configuration can be expressed in terms of the reference density ρR by

M(SR)=PRρRdV,M(BR)=RRρRdV,

si204_e  (4.4)

where dV is an element of volume in the reference configuration (refer to Section 3.6).

The densities ρ and ρR, defined in (4.1) and (4.3), respectively, both have material, Lagrangian, and Eulerian descriptions:

ρ=ˉρ(Y,t)=ˉρ(κ1(X),t)=ˆρ(X,t)=ˆρ(χ1(x,t),t)=˜ρ(x,t),ρR=ˉρR(Y)=ˉρR(κ1(X))=ˆρR(X)=ˆρR(χ1(x,t))=˜ρR(x,t).

si205_e

Note that since the material and Lagrangian descriptions of ρR have no time dependence,

˙ρR=tˉρR(Y)=tˆρR(X)=0.

si206_e  (4.5)

4.2 Forces and moments, linear and angular momentum

Again, consider the part Ssi207_e of body Bsi208_e that occupies region Psi209_e bounded by surface Psi210_e in the present configuration at time t. We denote the resultant external force acting on part Ssi211_e of the body by f:

f=f(S,t),the resultant external force acting on partSat timet,

si212_e

and the resultant external moment about the origin (point 0) by Mo:

Mo=Mo(S,t,0),the resultant external moment acting on partSat timetabout the origin0.

si213_e

We assume that the resultant external force f acting on part Ssi214_e at time t is composed of a body force fb and a contact force fc:

f=fb+fc.

si215_e  (4.6)

Additionally, we make the following smoothness assumptions:

fb=fb(S,t)=Pbρdv,fc=fc(S,t)=Ptda,

si216_e  (4.7)

where

b=˜b(x,t),thespecificbodyforce,i.e.,the body force per unit mass

si217_e

and

t=˜t(x,t,geometry of surface),thetraction,i.e.,the contact force per unit area of the present configuration.

si218_e

The geometry of the surface includes orientation, curvature, etc. The smoothness assumptions demand that b and t are bounded, continuous functions of space and time. We have already invoked a smoothness assumption on mass in the previous section, namely,

M(S,t)=Pρdv,

si219_e

where density ρ is a bounded, continuous function of space and time. With this smoothness assumption on mass, we have excluded point, line, and surface masses; our smoothness assumptions on forces exclude concentrated forces.

We assume there are no distributed moment fields and no concentrated moments (we later relax this assumption in Chapter 9 to accommodate the effects of electro-magnetism), so the resultant external moment Mo on part Ssi220_e at time t about the origin 0 comes from only t and b:

Mo(S,t,0)=P(x0)×bρdv+P(x0)×tda.

si221_e  (4.8)

Because of our smoothness assumption on mass, the linear momentum Lsi222_e of part Ssi223_e at time t and the angular momentum Ho of part Ssi224_e about the origin at time t are also smooth:

=(S,t)=Pvρdv,Ho=Ho(S,t,0)=P(x0)×vρdv.

si225_e  (4.9)

4.3 Equations of motion (mechanical conservation laws)

We postulate the following equations of motion:

 The mass of every subset Ssi226_e of the body remains constant throughout the motion, or, equivalently, the rate of change of the mass of Ssi227_e is zero (conservation of mass).

 The rate of change of linear momentum of Ssi228_e is equal to the resultant force acting on Ssi229_e (balance of linear momentum).

 The rate of change of angular momentum of Ssi230_e about the origin is equal to the resultant moment acting on Ssi231_e about the origin (balance of angular momentum).

Mathematically, these equations of motion can be expressed in material form as

ddtM(S,t)=0orM(S)=independent oft=M(SR),

si232_e  (4.10a)

ddt(S,t)=f(S,t),

si233_e  (4.10b)

ddtHo(S,t,0)=Mo(S,t,0)

si234_e  (4.10c)

for arbitrary part Ssi235_e of body Bsi236_e, and all time t. We emphasize that the equations of motion (4.10a)(4.10c) are global; i.e., they are valid not only for the body as a whole, but also for every arbitrary subset of the body. (The reader is already familiar with this requirement: as an example, in a static truss, not only is the entire structure in equilibrium, but so is each joint and each member.)

The smoothness assumptions discussed in Sections 4.1 and 4.2 can then be used to express the equations of motion (4.10a)1, (4.10b), and (4.10c) in Eulerian integral form:

ddtPρdv=0,

si237_e  (4.11a)

ddtPvρdv=Pbρdv+Ptda,

si238_e  (4.11b)

ddtPx×vρdv=Px×bρdv+Px×tda

si239_e  (4.11c)

for arbitrary material volume Psi240_e in the present configuration Rsi241_e of body Bsi242_e, and all time t. To perform the integrations in (4.11a)(4.11c) over areas and volumes in the present configuration, the functions ρ, v, b, and t must be in their Eulerian forms, i.e., functions of the independent variables x and t.

4.4 The first law of thermodynamics (conservation of energy)

To complete the set of fundamental laws of thermomechanics, we now postulate the law of conservation of energy, also known as the first law of thermodynamics. In general:

The rate of change of the total energy of any part Ssi243_e of the body is equal to the rate of mechanical work generated by the resultant external force acting on Ssi244_e plus the rate ofall other energies that enter or leave Ssi245_e (such as heat energy, chemical energy, or electromagnetic energy).

In this chapter, we specialize to thermomechanical systems, so the only other energy entering or leaving part Ssi246_e is heat. (This restriction is later relaxed in Chapter 9 to accommodate electromagnetic sources of energy.) It follows, then, that the law of conservation of energy in material form is

ddtT(S,t)=R(S,t)+H(S,t)

si247_e  (4.12)

for arbitrary subset Ssi248_e of body Bsi249_e and all time t, where

T=T(S,t)si250_e, the total energy of part Ssi251_e at time t,

R=R(S,t)si252_e, the rate of work done on part Ssi253_e at time t by the resultant external force f,

H=H(S,t)si254_e, the rate of heat energy entering part Ssi255_e at time t.

We emphasize that (4.12) is valid not only for the body as a whole, but also for every subset of the body.

The assumption (4.6) that the resultant external force f on Ssi256_e is the sum of a body force fb and a contact force fc implies that

R=Rb+Rc,

si257_e  (4.13)

and the smoothness assumptions on fb and fc (refer to (4.7)) further imply that

Rb(S,t)=Pb·vρdv(the rate of work of body forces),

si258_e  (4.14a)

Rc(S,t)=Pt·vda(the rate of work of contact forces),

si259_e  (4.14b)

where Psi260_e is the region of E3si261_e occupied by part Ssi262_e in the present configuration at time t, with boundary Psi263_e.

Our smoothness assumption (4.2)1 on mass implies that the energy of motion, or kinetic energy, of part Ssi264_e at time t is given by

K(S,t)=P12v·vρdv.

si265_e  (4.15)

We assume the existence of an internal energy E of part Ssi266_e of the body, such that

T(total energy)=K+E(kinetic energy+internal energy),

si267_e  (4.16)

and further make the smoothness assumption

E(S,t)=Pερdv,

si268_e  (4.17)

where

ε=˜ε(x,t),thespecific internal energy,i.e., the internal energy per unit mass,

si269_e

is a bounded, continuous function (so we have excluded point, line, and surface concentrations of internal energy).

We assume that the rate of heat energy H entering part Ssi270_e at time t is composed of two parts, that entering throughout the volume Psi271_e and that flowing through the surface Psi272_e:

H(S,t)=PrρdvPhda,

si273_e  (4.18)

where

r=˜r(x,t),thespecificheatsupplyrate,i.e.,the heat energy absorbed per unit mass per unit time,

si274_e

and

h=˜h(x,t,geometry of suface),theheatfluxrate,i.e.,the heat flowoutofPper unit area of the present configuration per unit time.

si275_e

Physically, (4.18) represents heat transfer due to radiation (first term) and conduction (second term). Again, we make the smoothness assumption that r and h are bounded, continuous functions. The minus sign appears in (4.18) because of the sign convention that h is positive when heat flows out of Psi276_e, i.e., through the surface Psi277_e in the direction of its outward unit normal n.

Using (4.13)(4.18), we can express the first law of thermodynamics (4.12) in Eulerian integral form:

ddtP12v·vρdv+ddtPερdv=Pb·vρdv+Pt·vda+PrρdvPhda

si278_e  (4.19)

for arbitrary material volume Psi279_e in the present configuration Rsi280_e of the body for all time t.

4.5 The transport and localization theorems

In this section, we present the transport theorem and the localization theorem. As will soon be evident, these are essential tools for rigorously deriving local (or pointwise) statements of the integral conservation laws developed in the previous two sections.

4.5.1 The transport theorem

The transport theorem allows us to take the time derivative of a volume integral whose region of integration Psi281_e changes with time. It is the three-dimensional analog of Leibniz's rule:

ddtb(t)a(t)f(x,t)dx=b(t)a(t)f(x,t)tdx+f(b(t),t)db(t)dtf(a(t),t)da(t)dt.

si282_e

Let Ssi283_e be an arbitrary part (or subset) of the body Bsi284_e that occupies a region PRsi285_e, with closed boundary PRsi286_e, in a fixed reference configuration, and occupies region Psi287_e, with closed boundary Psi288_e, in the present configuration at time t. Let ɸ be any continuous (in space and time) scalar-, vector-, or tensor-valued function, with the representations

ɸ=˜ɸ(x,t)=˜ɸ(χ(X,t),t)=ˆɸ(X,t).

si289_e

Then

ddtP˜ɸ(x,t)dv=P(˙ɸ+ɸdivv)dv.

si290_e  (4.20)

It can be shown that an alternative form of (4.20) is

ddtP˜ɸ(x,t)dv=Pɸdv+Pɸv·nda,

si291_e  (4.21)

where n is the outward unit normal to the surface Psi292_e. Refer to Section 3.2 for the definitions of ˙ɸsi293_e and ɸ′. (Note that the function ˜ɸ(x,t)si294_e must be continuous for ˙ɸsi295_e and ɸ′ to make sense.)

Exercises

1. Prove Leibniz's rule

ddtb(t)a(t)˜f(x,t)dx=b(t)a(t)˜f(x,t)tdx+˜f(b(t),t)db(t)dt˜f(a(t),t)da(t)dt

si296_e

as the one-dimensional analog of the three-dimensional transport theorem. (Hint: First, define a one-dimensional motion x = χ(X, t), and define the limits A and B such that χ(A, t) = a(t) and χ(B, t) = b(t). Proceed as in the proof of the three-dimensional transport theorem shown in Problem 4.1.)

Problem 4.1

Prove the transport theorem. That is, show that

ddtP˜ɸ(x,t)dv=P(˙ɸ+ɸdivv)dv.

si8_e

Solution

We begin by using dv = J dV (refer to (3.71)) and a change of independent variable from x to X to convert the Eulerian integration to a Lagrangian integration:

ddtP˜ɸ(x,t)dv=ddtPRˆɸ(X,t)ˆJ(X,t)dV.

si9_e

Since the region of integration PRsi10_e is fixed and independent of time t, it follows that

ddtPRˆɸ(X,t)ˆJ(X,t)dV=PRt[ˆɸ(X,t)ˆJ(X,t)]dV.

si11_e

The product rule, the definition (3.13) of the material derivative, and result (3.60)3 then imply that

PRt[ˆɸ(X,t)ˆJ(X,t)]dV=PR[tˆɸ(X,t)ˆJ(X,t)+ˆɸ(X,t)tˆJ(X,t)]dV=PR[tˆɸ(X,t)ˆJ(X,t)+ˆɸ(X,t)˙J]dV=PR[tˆɸ(X,t)ˆJ(X,t)+ˆɸ(X,t)ˆJ(X,t)divv]dV=PR[tˆɸ(X,t)+ˆɸ(X,t)divv]ˆJ(X,t)dV=PR[˙ɸ+ˆɸ(X,t)divv]ˆJ(X,t)dV.

si12_e

We once again use (3.71) and a change of independent variable from X to x, this time to convert the Lagrangian integration back to an Eulerian integration:

PR[˙ɸ+ˆɸ(X,t)divv]ˆJ(X,t)dV=P(˙ɸ+˜ɸ(x,t)divv)dv.

si13_e

Thus, we conclude that

ddtP(˜ɸ(x,t)dv)=P(˙ɸ+ɸdivv)dv.

si14_e

2. Using the definition of the material derivative and the divergence theorem, show that

P(˙ɸ+ɸdivv)dv=Pɸdv+Pɸv·nda.

si297_e

Then show that this result implies that

ddtP˜ɸ(x,t)dv=Pɸdv+Pɸv·nda,

si298_e

which is an alternative form of the transport theorem.

4.5.2 The localization theorem

If ɸ is a continuous scalar- or tensor-valued field in Rsi299_e and

Pɸdv=0

si300_e  (4.22)

for any part PRsi301_e, then it is necessary and sufficient that

ɸ=0

si302_e  (4.23)

in Rsi303_e. Said differently, if (4.22) holds for any arbitrary subset of the body, then the integrand ɸ vanishes everywhere throughout the body, and vice versa.

Problem 4.2

Prove the localization theorem.

Solution

That (4.23) implies (4.22) is trivial. To show that (4.22) implies (4.23), we first recall from real analysis the definition of continuity:

A function ɸ(x, t) is continuous in a region Rsi304_e if for every xR si305_e and every ε > 0 there exists a δ > 0 such that

|x0x|<δ|ɸ(x,t)ɸ(0x,t)|<ε.

si306_e

In what follows, we verify that (4.22) implies (4.23) using this definition of continuity and proof by contrapositive.

Suppose not (4.23), i.e., there is a point 0xRsi307_e at which ɸ0 = ɸ(0x, t) ≠ 0. Assume first that ɸ0 > 0. Since ɸ is continuous on Rsi308_e, we are free to select ε = ɸ0/2, and there must exist a δ > 0 such that

|x0x|<δ|ɸ(x,t)ɸ0|<ɸ02.

si309_e  (a)

Let Pδsi310_e be the region of all xRsi311_e such that |x0x| < δ, and let Vδ be the volume of this region, so

Vδ=Pδdv>0.

si312_e

From (a) it follows that

ɸ(x,t)>ɸ02inPδPδɸdv>ɸ02Pδdv=12ɸ0Vδ>0.

si313_e  (b)

Now assume ɸ0 < 0. By continuity of ɸ, there exists a δ > 0 such that

|x0x|<δ|ɸ(x,t)ɸ0|<ɸ02ɸ(x,t)<ɸ02inPδ,

si314_e

and therefore

Pδɸdv<ɸ02Pδdv=12ɸ0Vδ<0

si315_e  (c)

From (b) and (c) we see that, for either ɸ0 > 0 or ɸ0 < 0, we can find a PδRsi316_e for which (4.22) is not satisfied. Hence, we have shown

not(4.23)not(4.22),or equivalently,(4.22)(4.23).

si317_e

4.6 Cauchy stress tensor, heat flux vector

Recall from Section 4.2 that the traction t acting on the surface Psi318_e is, in general, a function of position x, time t, and the geometry of the surface, i.e.,

t=˜t(x,t,geometry of surface).

si319_e  (4.24)

The geometry of the surface includes orientation, curvature, and so on. We now restrict our attention to a particular type of contact force, namely, one such that the traction t is the same for all like-oriented surfaces with a common tangent plane at x and t. Therefore, the dependence of t on the geometry of the surface Psi320_e is only through the outward unit normal n, so (4.24) becomes

t=˜t(x,t,n).

si321_e  (4.25)

(An example of a traction that is not of the form (4.25) is surface tension, which depends on a measure of the curvature of the surface.)

We further assume that t is a continuous function of x, t, and n. Then, as a consequence of the conservation laws of mass (4.11a) and linear momentum (4.11b), and our assumptions that ¨xsi322_e, b, and ρ are bounded, it can be shown (refer to Problem 4.3) that the dependence of t on n in (4.25) is in fact linear, i.e.,

˜t(x,t,n)=ˆT(x,t)nort=Tn.

si323_e  (4.26)

The tensor T in (4.26) is called the Cauchy stress tensor. Note that the components of the Cauchy stress T are defined by

Tij=t(ej)·ei,

si324_e  (4.27)

where t(ej) is the traction on a surface whose unit normal is n = ej. This definition, along with result (4.26), implies that

Tij=ei·Tej,

si325_e  (4.28)

which is consistent with definition (2.29) of the Cartesian components of a tensor.

Problem 4.3

Prove that t = Tn.

Solution

Consider an arbitrary part Ssi15_e of body Bsi16_e that occupies a region Psi17_e in the present configuration at time t. Let Psi18_e be divided into two regions, P1si19_e and P2si20_e, separated by a surface σ (see Figure 4.1). Define Psi21_e and P′′si22_e so that P=PP′′si23_e. Then, P=P1P2si24_e, P=PP′′si25_e, P1=Pσsi26_e, P2=P′′σsi27_e.

f04-01-9780123946003
Figure 4.1 Schematic illustrating the division of Psi1_e into two regions, P1si2_e and P2si3_e, separated by a surface σ.

Applying balance of linear momentum to the three regions Psi28_e, P1si29_e, and P2si30_e, we obtain

ddtP˙xρdv=Pbρdv+Pt(n)da,

si31_e  (a)

ddtP1˙xρdv=P1bρdv+Pσt(n)da,

si32_e  (b)

ddtP2˙xρdv=P2bρdv+Pσt(n)da.

si33_e  (c)

Adding (b) and (c) then subtracting (a) gives

σ[t(n)+t(n)]da=0.

si34_e  (d)

In (d) we have noted that the outward normal of σ, when considered as a part of the boundary of P1si35_e, is directly opposed to the outward normal of σ when considered as a part of the boundary of P2si36_e. Assuming the traction is a continuous function of x and n, so that we can use the two-dimensional localization theorem, (d) implies

t(n)=t(n).

si37_e  (e)

(Traction vectors on opposite sides of the same surface are equal in magnitude, and opposite in direction.)

Consider now a family of similar tetrahedra Tsi38_e with heights hp and a common vertex at some point ox (see Figure 4.2). Sides i are perpendicular to the xi directions and have outward normals −ei. The remaining side has outward normal on. From geometry, areas Ssi39_e, S1si40_e, S2si41_e, and S3si42_e are related by

Si=S(on·ei)=Sni.

si43_e  (f)

The volumes of the tetrahedra are

V=13hpS.

si44_e  (g)

It is assumed that each of the tetrahedra lies completely within the region Rsi45_e occupied by Bsi46_e in the present configuration. By virtue of the transport theorem (4.20) and the local form of conservation of mass (presented later in Section 4.8), (a) becomes

P¨xρdv=Pbρdv+Pt(n)da.

si47_e  (h)

Applying (h) to a tetrahedron Tsi48_e gives

T(x¨b)ρdv=ΣSit(ei)da+St(on)da.

si49_e  (i)

With use of (e), (i) becomes

T(x¨b)ρdv=St(on)daΣSit(ei)da.

si50_e  (j)

Recall that the specific body force b and density ρ are assumed bounded throughout Rsi51_e. The acceleration ¨xsi52_e is also assumed bounded. Then,

|T(¨xb)ρdv|T|¨xb|ρdv(by a theorem of analysis)=TKdv(by boundedness)=KTdv=KShp3(using(g)),

si53_e  (k)

where K is some number. Assuming that t is a continuous function of x and t, the mean value theorem and (f) imply that

St(on)daΣSit(ei)da=t*(on)St*(ei)Si=[t*(on)t*(ei)ni]S,

si54_e  (l)

where t*(on) and t*(ei) stand for some specific interior values of the traction vectors on the faces Ssi55_e and Sisi56_e, respectively. From (j) to (l),

13KShp|T|¨xb|ρdv|=|St(on)daΣSit(ei)da|=|t*(on)t*(ei)ni|S,

si57_e

so

|t*(on)t*(ei)ni|13Khp.

si58_e

As we consider smaller and smaller tetrahedra, hp → 0, and

limhp0|t*(on)t*(ei)ni|=0,

si59_e  (m)

where t*(on) and t*(ei) are evaluated at the point ox, which is the common vertex of the family of tetrahedra. From (m), we see that in the limit hp → 0, t*(on) − t*(ei)ni must be the zero vector, i.e.,

t*(on)=t*(ei)oni,

si60_e  (n)

where t*(ei)o denotes the value of t*(ei) at the point ox. Since (n) must hold at any point ox and corresponding to any direction on, without ambiguity, we may delete the star, replace ox and on by x and n, and write

t(n)=t(ei)ni.

si61_e  (o)

We define Tki by

Tki=t(ei)·ek,

si62_e

and denote the components of t(n) by

tk=t·ek.

si63_e

Taking the inner product of (o) with the unit vector ek thus gives

tk=t(ei)ni·ek=Tkini.

si64_e  (p)

Equation (p) is the Cartesian component form of

t=Tnort(x,t,n)=T(x,t)n.

si65_e

f04-02-9780123946003
Figure 4.2 A tetrahedron of height hp. Sides S1si4_e, S2si5_e, and S3si6_e have outward unit normals −e1, −e2, and −e3, respectively, while side S has outward unit normal on.

Recall from Section 4.4 that the heat flux h out of the surface Psi326_e is, in general, a function of position x, time t, and the geometry of the surface, i.e.,

h=˜h(x,t,geometry of surface).

si327_e  (4.29)

We restrict our attention to a particular type of heat flow through the boundary Psi328_e, namely, one such that the heat flux h is the same for all like-oriented surfaces with a common tangent plane at x and t, so (4.29) becomes

h=˜h(x,t,n).

si329_e  (4.30)

We further assume that h is a continuous function of x, t, and n. Then, as a consequence of the conservation laws of mass (4.11a), linear momentum (4.11b), and energy (4.19), and our boundedness assumptions, it can be shown (refer to Problem 4.4) that the dependence of h on n in (4.30) must be linear, i.e.,

˜h(x,t,n)=˜q(x,t)·norh=q·n.

si330_e  (4.31)

The vector q in (4.31) is called the heat flux vector.

Problem 4.4

Show that h = q · n.

Solution

Consider once again the regions Psi66_e, P1si67_e, and P2si68_e shown in Figure 4.1. Applying the first law of thermodynamics to these three regions, we have

ddtP(˙x·˙x2+ε)ρdv=P(b·v+r)ρdv+P[t(n)·vh(n)]da,

si69_e  (a)

ddtP1(˙x·˙x2+ε)ρdv=P1(b·v+r)ρdv+Pσ[t(n)·vh(n)]da,

si70_e  (b)

ddtP2(˙x·˙x2+ε)ρdv=P2(b·v+r)ρdv+Pσ[t(n)·vh(n)]da,

si71_e  (c)

Adding (b) and (c) then subtracting (a) gives

σ[t(n)+t(n)]·vda+σ[h(n)+h(n)]da=0.

si72_e  (d)

In (d) we have noted that the outward normal of σ when considered as a part of P1si73_e is directly opposed to the outward normal of σ when considered as a part of P2si74_e. The result

t(n)=t(n)

si75_e

from Problem 4.3 implies that (d) becomes

σ[h(n)+h(n)]da=0.

si76_e

Assuming continuity of h, it then follows that

h(n)=h(n).

si77_e  (e)

(The heat leaving one part equals the heat entering the other part.)

By virtue of the transport theorem (4.20) and the local form of conservation of mass (presented later in Section 4.8), (a) reduces to

P(v·˙v+˙ε)ρdv=P(b·v+r)ρdv+P[t(n)·vh(n)]da.

si78_e

Applying this to the tetrahedron Tsi79_e in Figure 4.2 gives

T(v·˙v+˙εb·vr)ρdv=ΣSi[t(ei)·vh(ei)]da+S[t(on)·vh(on)]da,

si80_e

or, using (e),

T(v·˙v+˙εb·vr)ρdv=ΣSi[t(ei)·v+h(ei)]da+S[t(on)·vh(on)]da.

si81_e  (f)

We assume that ε, ˙εsi82_e, v, ˙vsi83_e, and r are bounded, so

|T(v·˙v+˙εb·vr)ρdv|MShp3

si84_e  (g)

for some number M. Therefore, from (f), (g), and the mean value theorem, we have

13MShp|ΣSi[t(ei)·v+h(ei)]da+S[t(on)·vh(on)]da|=|[t*(on)·v*h*(on)]S+[t*(ei)·v*(i)+h*(ei)]Si|=|t*(on)·v*t*(ei)ni·v*(i)h*(on)+h*(ei)ni|S,

si85_e

where t*(on), v*, and h*(on) stand for some specific interior values on face Ssi86_e, and t*(ei), v*(i), and h*(ei) stand for some specific values on faces Sisi87_e. Then

limhp0|t*(on)·v*t*(ei)ni·v*(i)h*(on)+h*(ei)ni|=0,

si88_e

so in the limit,

t*(on)·v*t*(ei)ni·v*(i)h*(on)+h*(ei)ni=0.

si89_e  (h)

In the limit, v* = v*(i) = v(ox, t), and so (h) becomes

[t*(on)t*(ei)ni]·v*h*(on)+h*(ei)ni=0.

si90_e  (i)

Recall from Problem 4.3 that

t*(on)=t*(ei)oni.

si91_e

As a consequence, (i) becomes

h*(on)=h*(ei)niorh(n)+h(ei)ni.

si92_e  (j)

We define qi by

qi=h(ei)

si93_e

so that (j) is written as

h(n)=qini.

si94_e  (k)

Equation (k) is the Cartesian component form of

h=q·norh(x,t,n)=q(x,t)·n.

si95_e

4.7 The energy theorem and stress power

Recall from Section 4.3 the Eulerian representations of the mechanical integral conservation laws:

ddtPρdv=0,

si331_e  (4.32a)

ddtPvρdv=Pbρdv+PTnda,

si332_e  (4.32b)

ddtPx×vρdv=Px×bρdv+Px×Tda,

si333_e  (4.32c)

where we have used t = Tn (result (4.26)). As a consequence of these mechanical conservation laws, it can be shown (refer to Problem 4.7) that the rate of work of all external forces acting on Psi334_e and its boundary Psi335_e minus the rate of increase of the kinetic energy is equal to the total stress power of Psi336_e:

Pb·vρdv+Pt·vdaddtP12v·vρdv+PT·Ddv.

si337_e  (4.33)

This result, called the energy theorem, is a mechanical result: it follows only from the mechanical laws (4.32a)(4.32c) and does not depend on conservation of energy. The scalar quantity T · D in (4.33) is called the stress power P, i.e.,

P=T·D.

si338_e  (4.34)

Problem 4.7

Verify that the Eulerian (or spatial) form of the energy theorem

Pb·vρdv+Pt·vdaddtP12v·vρdv=PT·Ddv

si122_e

is a consequence of the mechanical conservation laws of mass, linear momentum, and angular momentum.

Solution

To begin, we consider the second term on the left-hand side of the energy theorem:

Pt·vda=PTn·vda(result(4.26))=PTTv·nda(definition(2.13))=Pdiv(TTv)dv(divergence theorem(2.104)2)=P(T·gradv+v·divT)dv(result(2.99)4).

si123_e

Note that this result is consistent with (2.105). Also note that grad and div are the gradient and divergence calculated with respect to the present configuration. Next, considering the third term on the left-hand side of the energy theorem, we have

ddtP12v·vρdv=P(12.¯v·vρ+12v·vρdivv)dv(transport theorem(4.20))=P[ρ˙v·v+12v·v(˙ρ+ρdivv)]dv(product rule(3.22))=Pρ˙v·vdv(conservation of mass(4.36a)).

si124_e

After assembling these results, the left-hand side of the energy theorem becomes

P[(ρb+divTρ˙v)·v+T·gradv]dv.

si125_e

By virtue of the local form of conservation of linear momentum (4.36b), this further reduces to

PT·gradvdv.

si126_e

Then,

PT·gradvdv=PT·Ldv(definition(3.56))=PT·(D+W)dv(decomposition(3.57))=P(T·D+T·W)dv(property(2.6)3)=PT·Ddv(result(2.44)),

si127_e

where, in the last step, we have exploited the local form of conservation of angular momentum (4.36c), which demands that T is symmetric; recall that W is skew by construction. (Note that we were able to employ the distributive property (2.6)3 of the inner product in the next-to-last step since we demonstrated in Problem 2.35 that the set of all second-order tensors is an inner product space.) Thus, we conclude that

Pb·vρdv+Pt·vdaddtP12v·vρdv=PT·Ddv.

si128_e

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset